You are on page 1of 18

Water Air Soil Pollut (2017) 228:278

DOI 10.1007/s11270-017-3461-y

Synthesis of a Quaternized Beta


Cyclodextrin-Montmorillonite Composite and Its Adsorption
Capacity for Cr(VI), Methyl Orange, and p-Nitrophenol
Anrong Zeng & Anran Zeng

Received: 24 March 2017 / Accepted: 27 June 2017


# Springer International Publishing AG 2017

Abstract In this paper, quaternized β-cyclodextrin– process and could be spontaneous at given temperature
montmorillonite composite (QCD-MMT) was obtained range, except for Cr(VI), of which adsorption should be
and absorption properties of Cr(VI), methyl orange and at much higher temperature. Overall, QCD-MMT ex-
p-nitrophenol were studied. QCD was prepared by 2,3- hibited potential for practical applications in the treat-
epoxypropyltrimethylammonium chloride and β- ment of both metal ions and organic pollutants.
cyclodextrin (β-CD). QCD-MMT was obtained by re-
action between QCD and montmorillonite suspensions, Keywords Beta-Cyclodextrin . Montmorillonite .
which could be attributed to the montmorillonite cation Composite . Adsorption capacity
ion exchange properties. β-CD cavities of this compos-
ite were expected to capture organic molecules through
inclusion, while montmorillonite units acted as the ad-
1 Introduction
sorption sites for metals. QCD-MMT was characterized
by FT-IR, elemental analysis, XRD, SEM-EDX, and
With rapid development of industrialization and urban-
TGA. Adsorptions of Cr(VI), methyl orange, and p-
ization, a variety of hazardous molecules was used and
nitrophenol were highly dependent on adsorbent dose,
could be possibly discharged into the natural environ-
initial concentration, temperature, contact time, and pH.
ment. For example, synthetic dyes were commonly used
Adsorption kinetics of Cr(VI), methyl orange, and p-
in textile, cosmetics, and other industries (Yan et al.
nitrophenol followed the pseudo-second-order model.
2016). Dye containing wastewater could endanger hu-
Meanwhile, adsorption of Cr(VI) fit better in the
man health without effective treatment. Also, pollutants
Freundlich model, inferring a multilayer adsorption,
like p-nitrophenol has become one of the commonly
while the adsorption of methyl orange and p-
encountered pollutants in water pollution, and it could
nitrophenol fit better in Langmuir model, inferring a
be persistent, bio-accumulative, and highly toxic at very
monolayer adsorption. Thermodynamic analysis
low concentration (Shen et al. 2015). Moreover, metal-
showed that the adsorptions were all endothermic
bearing effluents originating from metal plating, mining
operations, nuclear power plant, and metallurgical and
A. Zeng (*) : A. Zeng
battery manufacturing industries, often contained many
College of Light–Textile Engineering, Liming Vocational kinds of heavy metal ions, such as lead, cobalt, nickel,
University, Quanzhou 362000, China cadmium, chromium, and mercury ions. Heavy metals
e-mail: anrongzeng@qq.com were non-biodegradable and can accumulate slowly in
A. Zeng : A. Zeng
the body of living creatures, causing various diseases
Applied Technology Center of Fujian Provincial Higher Education such as poisoning, nervous system damage, and cancer
for Practical Chemical Material, Quanzhou 362000, China (Vašák and Meloni, 2011; Zheng 2001). Overall, those
278 Page 2 of 17 Water Air Soil Pollut (2017) 228:278

pollutants could further lead to many serious environ- 2014), it was widely used as an adsorbent for water
mental problems, like water pollution and further seri- treatment, especially for heavy metal treatment.
ous public health issue. Therefore, effective treatments In this work, we prepared a β-CD/MMT composite
are becoming more and more urgent. (QCD-MMT) for adsorption of Cr(VI) ions (as the
A lot of water treatment methods have been reported. representative of heavy metal ions), methyl orange (as
For example, adsorption, ion exchange, and reverse the representative of dye compounds), and p-
osmosis have been applied to the removal of chromium nitrophenol (as the representative of phenols pollutants),
and mercury in paper and textile industries (Liu et al. in aqueous solution. In order to further investigate the
2016). Oxidation, membrane separation, and extraction interaction between QCD-MMT and different adsor-
were also developed to remove p-nitrophenol or its bates, the adsorption behavior was studied in terms of
analogues from wastewater (Shen et al. 2015). For dyes various parameters such as adsorption kinetics, adsorp-
containing wastewater treatment, methods like adsorp- tion isotherms, and thermodynamic models.
tion, oxidation process, precipitation, coagulation, and
electro-chemical treatment could be applied (Debnath
et al. 2017). Among these approaches, adsorption was 2 Materials and Methods
used most frequently. It was cost saving, efficient, and
promising to remove both organics and heavy metals 2.1 Materials
from wastewater (Chen et al. 2015; Debnath et al. 2017).
β-cyclodextrin (β-CD) has been found to be effective to β-CD was 96% pure and purchased from Aladdin. MMT
remove heavy metals (Liu et al. 2016). It is usually used was purchased from Zhejiang Fenghong New Material
as a kind of pharmaceutical excipients. β-CD has a Co. Ltd., industrial grade; 2,3-epoxypropyltrimethyl am-
torus-shaped cyclic oligosaccharide structure with abun- monium chloride (EPTAC) was purchased from Shang-
dant hydroxyl groups on both ends, and a hydrophobic hai Dibo Chemical Co. Ltd., analytical grade. Other
cavity. As a result, it shows good ability to capture reagents were purchased from Xilong Scientific, analyti-
organic molecules. However, the water-solubility of β- cal grade. All reagents were used without further
CD limits its application to some extent, and it is im- purification.
portant to improve its insolubility in water. One way Potassium dichromate, methyl orange, and p-
reported was to use cross-linkers to obtain β-CD poly- nitrophenol stock solutions were prepared via dissolving
mers. For example, EDTA-β-CD was employed in the in deionized water. Working solutions of metals and
adsorption of aqueous dyes and metal ions (Zhao, Repo, dyes were prepared by diluting the stock solutions.
Meng, et al., 2015b; Zhao, Repo, Yin, et al., 2015a). The The three adsorbates properties were summarized in
enhanced metal adsorption ability of β-CD was attrib- Table 1.
uted to EDTA’s functional groups of chelating metal
ions. Similar modifications were reported to improve 2.2 Synthesis of QCD-MMT Composite
the chitosan’s metal ion affinities as well, by using
EGTA, EDTA, and DTPA (Zhao, Repo, Sillanpää, The synthetic process of QCD-MMT was shown in Fig.
et al., 2015c; Zhao et al. 2013). Another way to increase 1. β-CD was dissolved in NaOH solution (pH = 12) in a
β-CD insolubility was to prepare inorganic-organic flask equipped with a mechanical stirrer, thermometer,
composite adsorbents, which has been applied to im- reflux cooler, and drop funnel. EPTAC aqueous solution
prove certain polymer’s adsorption performance. For (75% by weight) was added into the drop funnel and
instance, hydroxamic-acid-modified polyacrylamide/ further added dropwise into the flake. Mass ration of
Fe3O4 composite was prepared and showed affinity EPTAC and β-CD was 80/20. The reaction was con-
towards Cd(II), Pb(II), Co(II), and Ni(II) (Zhao et al. ducted at 60 °C for 3 h. Then the reaction system was
2014). cooled to room temperature to obtain QCD.
Montmorillonite (MMT), with an ideal molecular MMT was suspended in water with QCD. Mass ratio
formula of [Al2(Si4O10)(OH)2 ∙ XH2O], is present in a of MMT and β-CD was of 10/1. The reaction was
type of hydrous, porous aluminum silicate layer mineral. further conducted for another 24 h while sufficiently
MMT mines are abundant in nature, and their price is stirring at room temperature. The pH was adjusted to
quite low. In terms of its cheapness and features (Li et al. neutral with hydrochloric acid solution. The solid was
Water Air Soil Pollut (2017) 228:278 Page 3 of 17 278

Table 1 Adsorbates used in this paper

Adsorbates Chemical Formula

Potassium dichromate K2Cr2O7

Methyl orange

p-nitrophenol

further purified with deionized water, followed by dry- MMT’s thermal stability and organic content. Samples
ing in a vacuum oven of 60 °C. The final product was were heated to 800 °C at a heating rate of 10 °C/min in a
QCD-MMT composite. nitrogen flow of 40 ml/min.
The organic content was calculated as the percentage of
2.3 Characterization of QCD-MMT Composite organic mass with respect to the dried mass of MMT. The
dried mass of initial mass was obtained at 200 °C from
β-CD, MMT, and QCD-MMT FT-IR spectra were obtain- thermogravimetric curves in order to avoid the contribu-
ed using Bio-Rad FTS 6000 spectrometer. All samples tion of water, which could introduce error if samples were
were prepared as KBr pellets. Elemental analysis was not dried correctly. The mass of organic compounds was
determined using an Elementar Vario EL Cube analyzer, calculated from the mass loss registered between 200 and
CHN mode. X-Ray diffraction (XRD) data of MMT and 800 °C, so the organic content would be calculated as
QCD-MMT power samples were performed with Bruker following Eq. (1) (Ezquerro et al. 2015):
powder X-ray diffractometer D8 Advance, with scan range  
from 5° to 90° (2θ). The surface morphology and energy- mi −m f
Orangic content ¼  100 −%ΔmMMT ð1Þ
dispersive X-ray (EDX) spectroscopy were performed by mi
field emission scanning electron microscopy (FE-SEM,
Hitachi S4800). Thermogravimetric analysis (TGA) was where mi and mf were the mass of QCD-MMT at 200 and
also conducted on a TA instrument Q500 to assess QCD- 800 °C, respectively; %ΔmMMT was the mass loss

Fig. 1 Synthesis of QCD-MMT


composite
278 Page 4 of 17 Water Air Soil Pollut (2017) 228:278

observed in the same region of temperature for the unmod- 2.4.2 Adsorption Kinetics Analysis
ified sample. The decomposition of organics took place in
the range of 200 to 500 °C. The final temperature was To determine QCD-MMT’s adsorption kinetics of
chosen at 800 °C, since some overlap with the Cr(VI), methyl orange and p-nitrophenol, independent
dihydroxylation of silicate layers was observed at around bottles containing 30 mL samples and 0.05 g QCD-
500 °C. MMT, were used to get accurate results for each point
The ratio of quaternary ammonium groups of QCD- on the graph. The initial concentration of potassium
MMT was determined by N element ratio from element dichromate solution, methyl orange solution and p-
analysis, compared with the results of Cl element per- nitrophenol solution was 360, 250, and 25 mg/L, re-
centage from EDX. spectively. Adsorption experiments’ contact time varied
from 15 to 180 min. Adsorption was carried out at the
temperature of 30 °C.
2.4 Adsorption Experiments In order to investigate the adsorption mechanism of
three different adsorbates, two commonly used reaction
2.4.1 Adsorption Parameters’ Impact on QCD-MMT’s kinetic models viz., pseudo-first-order and pseudo-
Adsorption Performance second-order models were used (Xing and Wang
2016). The kinetic data were utilized to investigate the
All adsorption experiments were undertaken by mixing suitability of these models. The linear equations of
adsorbents with 30 mL potassium dichromate, methyl pseudo-first-order and pseudo-second-order models
orange, and p-nitrophenol solutions of different initial were expressed as Eqs. (3) and (4), respectively:
mass concentrations. The impacts of usage, initial con-
centration, temperature, contact time, and pH were qt ¼ qe1 −qe1 e−k 1 t ð3Þ
determined.
For each adsorption system, the residual in the reac-
tion mixture was analyzed by centrifuging the reaction t t 1
¼ þ ð4Þ
mixture, and then measuring the absorbance of the qt qe k 2 qe 2
supernatant at the wavelength corresponding to the
In Eq. (3), qe1 and qt were the adsorption capacity at
maximum absorbance of the sample. Concentrations in
equilibrium and at time t, respectively (mg/g); k1 was the
the reaction mixture were calculated from the calibration
rate constant of pseudo-first-order model (min−1). In Eq.
curve. Methyl orange and p-nitrophenol concentrations
(4), qe2 and qt were the adsorption capacity at equilibri-
were determined by UV-vis spectrometry (T6 model,
um and at time t; k2 (slope2/intercept) was the pseudo-
Beijing Purkinje General Instrument Co., Ltd.) at the
second-order rate constant (g/mg/min).
maximum absorbance of dyes. Metal concentrations
were analyzed by an atomic absorption spectroscopy
(A3 model, Beijing Purkinje General Instrument Co., 2.4.3 Adsorption Isotherms Analysis
Ltd.).
The amount of adsorbate adsorbed onto QCD-MMT, Isothermal studies were conducted with QCD-MMT
qe (mg/g), was calculated by the following mass balance usage of 0.05 g and 30 mL adsorbates solutions by
relationship: shaking the reaction mixture for equilibrium time. The
adsorption temperature was of 303.15, 313.15, and
323.15 K, respectively.
V The Langmuir model is valid for monolayer adsorp-
qe ¼ ðC 0 −C e Þ ð2Þ
W tion onto a surface with a finite number of identical sites.
The Langmuir equation was as follows (Xing and
Wang, 2016):
where C0 and Ce were initial and equilibrium liquid-
qm K L C e
phase concentration of waste, respectively (mg/L), V qe ¼ ð5Þ
1 þ K L Ce
was the volume of the solution (L), and W was the
weight of the QCD-MMT used (g). where Ce and qe were equilibrium concentration (mg/L)
Water Air Soil Pollut (2017) 228:278 Page 5 of 17 278

and the amount of adsorbates adsorbed on QCD-MMT that β-CD was formed by glucopyranose units
(mg/g), respectively. The constant KL was related to the through α-1,4-glycosidic bond (Yuan et al. 2015).
energy of adsorption and the constant qm represents the The peak of 1414 cm−1 was one of the typical
maximum amount of adsorbates absorbed per unit peaks corresponding to the deformation vibration
weight of QCD-MMT. of the O-H bonds in the primary and secondary
The Freundlich model can be applied to non-ideal hydroxy groups of β-CD, and also C-H bending
sorption on heterogeneous surfaces as well as multilayer vibrations (Zhao et al. 2009). It could be inferred
sorption and was expressed by the following equation that the structure characteristics of β-CD were
(Xing and Wang, 2016): maintained. Moreover, compared MMT with
1
QCD-MMT, the peaks at 521 and 465 cm−1 could
qe ¼ K F C en ð6Þ be attributed to the Si-O-Al stretching vibration
and Si-O bending vibration. Therefore, it could
where KF and n were the Freundlich constants. KF and n
be concluded that the modification of β-CD and
are indicators of adsorption capacity and adsorption
MMT was successful.
intensity of the QCD-MMT, respectively.

2.4.4 Adsorption Thermodynamics Analysis


3.1.2 Element Analysis
The thermodynamic parameters (ΔH , ΔS , and ΔG )
0 0 0
The results of quantitative elemental analysis were
for three different adsorbates onto QCD-MMT can be
presented in Table 2. Data showed the C and N
calculated from the temperature-dependent adsorption
element content of QCD-MMT was increased. The
isotherms. The equation was as follows (Li et al. 2013;
calculated quaternary ammonium group percentage
Xing and Wang, 2016):
was 8.22 wt.% basing on N element content. This
ΔS 0 ΔH 0 proved the final product was quaternized success-
lnK 0 ¼ − ð7Þ fully. This inference was echoed by EDX. EDX-Cl
R RT
element content was also given in Table 2, and the
where R is the universal gas constant (8.314 kJ/(mol
calculated quaternary ammonium group percentage
K)); T is the temperature in Kelvin; K0 is the adsorption
was 13.02 wt.%. EDX spectrum was recorded to
equilibrium constant and can be calculated by plotting ln
identify the expected compositions on a selected
Kd versus Ce and extrapolating Ce to zero, where Kd was
rectangle area of MMT and QCD-MMT sediment.
defined as qe divided by Ce.
The difference between two results might be ex-
plained by uneven quaternary ammonium group
distribution. There might be more quaternary am-
3 Results and Discussion monium groups on QCD-MMT’s surface. In this
paper, results based on element analysis were de-
3.1 Characterizations of QCD-MMT termined as the quaternary ammonium group
percentages.
3.1.1 FT-IR Spectroscopy

Figure 2 showed the FT-IR spectra of raw β-CD, 3.1.3 X ray Diffraction (XRD)
raw MMT, and QCD-MMT composite. Compared
with MMT, the peaks at 2928 cm−1 of QCD-MMT Figure 3 displayed the XRD pattern of raw MMT and
could be attributed to the methyl groups’ asym- QCD-MMT. It showed that raw MMT has a d (001)-
metric stretching vibration (Zhao et al. 2009). The spacing of 1.335 nm, basing on Bragg Eq.
peaks at 1487 cm−1 of QCD-MMT indicated the (2)d001sinθ = λ (λ = 0.154 nm) calculation. After mod-
existence of −CH2-N+, compared to β-CD, and ification, the QCD-MMT’s d001-spacing was enlarged to
could be assigned to NH3+. Besides, a characteris- 1.409 nm. It could be inferred that the modification may
tic absorption peak of α-type glycosidic bond was occur in the MMT interlayer. This might also benefit
found at 855 cm−1 of QCD-MMT, which indicates QCD-MMT’s adsorption capacity.
278 Page 6 of 17 Water Air Soil Pollut (2017) 228:278

Fig. 2 FT-IR spectra of MMT, β-


CD, and QCD-MMT

3.1.4 SEM-EDX Analysis identify the expected composition on a selected


rectangle area, we took EDX as supplement data
SEM provided information about morphology changes for element analysis.
before and after modification. Images were collected in
Fig. 4a, b. In the pristine state, MMT showed a rough 3.1.5 Thermogravimetric Analysis
surface where large lamellas were observed. QCD-
MMT’s surface was smooth after modifications with TGA and derivative thermogravimetric (DTG)
QCD. curves were shown in Figs. 6 and 7, respectively,
Simultaneously to SEM images, EDX spectros- and were also summarized in Table 4. The DTG
copy analysis was also carried out to determine the curves provided information about different steps
quaternary ammonium amount. Results were of decompositions. The thermal decomposition of
shown in Fig. 5 and Table 3. Cl element content MMT could be divided into four different regions
was increased from 0.09 wt.% to 3.14 wt.%. Cl (Ezquerro et al. 2015; Xie et al. 2014). The first
ion was the counter ion of quaternary ammonium weight loss region was desorption of water and
in QCD-MMT and indicated that the MMT had physisorbed species, which happened below
been quaternized successfully (Rosa et al. 2008; 180 °C. The second region was located in the
Sowmya and Meenakshi, 2013; Yao et al. 2014). range from 200 to 500 °C and involves the de-
However, since EDX only analyzes the surface and composition of organic species. The third region

Table 2 MMT and QCD-MMT element analysis

Sample Element content (wt.%) Total quaternary ammonium group wt.% Quaternary ammonium group wt.% (EDX result)

N C H

MMT 0.56 0.95 1.827 / /


QCD-MMT 1.32 11.19 2.709 8.22a 13.02b
a
Calculated from nitrogen content of element analysis: quaternary ammonium group wt.% = (1.32–0.56)/MN * MEPTAC, where MN stands
for molar mass of N element, MEPTAC stands for molar mass of EPTAC
b
Calculated from chlorine content of EDX analysis: quaternary ammonium group wt.% = (3.14–0.09)/MCl * MEPTAC,xxx where MCl stands
for molar mass of Cl element, MEPTAC stands for molar mass of EPTAC
Water Air Soil Pollut (2017) 228:278 Page 7 of 17 278

Fig. 3 XRD patterns of MMT


and QCD-MMT

was from 500 to 700 °C, when aluminosilicate around 9.10 wt%. Those results were also given in
layers’ dehydroxylation took place. Finally, in the Table 4.
last step between 700 and 1000 °C, the evolution
of products associated with organic carbonaceous
residue occurs. 3.2 Cr(VI), Methyl Orange and p-Nitrophenol
The position of these steps in the derivative Adsorption Experiments
mass curves is referred as the temperature of max-
imum decomposition rate in Table 4 (T1, T2, and 3.2.1 Influence of QCD-MMT Usage
T3). T1 corresponds to desorption of water and
other physisorbed species, T2 refers to the decom- The influence of adsorbent usage was studied, and
position step of organic species, and T3 refers to the results were shown in Fig. 8, where qe represent-
the aluminosilicate layers’ dihydroxylation. The ed the sorption amount of Cr(VI), methyl orange and
decomposition of carbonaceous residue was not p-nitrophenol on QCD-MMT (mg/g) in equilibrium
registered in the curves below the maximum tested time. At the usage of 0.05 g, the adsorption capacity
temperature of 800 °C. for Cr(VI), methyl orange, and p-nitrophenol was
DTG of MMT showed only two peaks (T1 and 51.98, 25.00, and 24.56 mg/g, respectively. Adding
T3). The first one was below 100 °C (T1), and the more adsorbent will decrease the adsorption capacity.
mass loss was associated to the evaporation of It could be also noticed that the difference between
water adsorbed. The second decomposition step initial concentration and Ce did not fluctuate a lot (Ce
occurred in the range 550–750 °C (T3) and could values of Cr(VI), methyl orange, and p-nitrophenol
be attributed to the dihydroxylation of the layer shown in Fig. 8). This could be explained by that the
crystal lattice. QCD-MMT showed one additional adsorption was driven by adsorbates’ concentration
peak at around 304 °C, which was associated to gradient (Xing and Wang, 2016). Given that concen-
the decomposition of organic compound introduced tration gradients in each adsorption were similar and
to MMT. the total adsorbates adsorbed did not change a lot, the
Organic content of QCD-MMT was calculated adsorption capacity would be decreased when QCD-
basing on TGA result, which was 16.25 wt.%. MMT usage increased. Also, the aggregation and
Considering element analysis results, CD content competition of QCD-MMT adsorbent might also
was around 8.03 wt.%, while QCD content was have impacts the adsorption behaviors.
278 Page 8 of 17 Water Air Soil Pollut (2017) 228:278

g at 50 °C, of which initial concentration was


200 mg/L. The increasing initial p-nitrophenol
concentrations could not always increase adsorp-
tion. On the contrary, the adsorption was slightly
lower. It was possible that high concentrations
resulted in competition between p-nitrophenol and
then slowed down the adsorption. Also, high con-
centration of p-nitrophenol might occupy and
block the QCD-MMT inclusion sites at the sur-
face, preventing further p-nitrophenol dispersion.

3.2.3 Influence of Contact Time

In order to determine the adsorption equilibration


time of different adsorbates and also to investigate
the kinetics of adsorption process, the adsorption
capacity as a function of contact time was studied.
Three different samples showed similar trends.
As adsorption time changed from 15 to 180 min,
the adsorption capacity first increased steadily then
afterwards no significant change observed. For
Cr(VI) and methyl orange, near equilibrium was
achieved in around 1.5 h, while for p-nitrophenol
in around 2.5 h. This was probably due to larger
surface area of QCD-MMT being available at the
beginning for adsorptions. As time progressed, the
surface coverage of the adsorbent was high and
then no adsorption took place. Figure 10a–c
showed the effect of time on adsorption. The equi-
librium adsorption capacities of Cr(VI), methyl
orange, and p-nitrophenol were around 43, 50,
and 6.5 mg/g, respectively.
Fig. 4 SEM images of MMT (a) and QCD-MMT (b)
3.2.4 Influence of pH

3.2.2 Influence of Initial Concentration Impact of pH was shown in Fig. 11. The adsorp-
tion capacities of Cr(VI) at different pH environ-
Figure 9 showed the influence of the initial concen- ment ranged from 34.80 to 51.98 mg/g. The ad-
trations on QCD-MMT’s sorption levels, respec- sorption capacities were higher at acidic and neu-
tively. The adsorption capacities of Cr(VI) in- tral condition. Maximum adsorption capacity of
creased as the initial concentrations increased. Cr(VI) reached the maximum at pH of 7. At lower
Methyl orange and p-nitrophenol showed similar pH, HCrO4− was the major form of Cr(VI). It is
trends. When the initial methyl orange concentra- possible that the quaternary ammonium groups on
tion was up to 250 mg/L at different temperatures the adsorbents were easily protonated and positive-
of 30, 40 and 50 °C, the adsorption capacity was ly charged, which facilitated the approach of the
53.217, 58.784, and 64.881 mg/g, respectively. For negatively charged HCrO4− due to the electrostatic
p-nitrophenol, when the initial concentration was interaction (Liu et al. 2016). Also, at lower pH,
increased, the adsorption capacity was largely en- the counter anion Cl− of quaternary ammonium
hanced. The maximum adsorption was 51.212 mg/ groups might be displaced by HCrO 4 − (Wang
Water Air Soil Pollut (2017) 228:278 Page 9 of 17 278

Fig. 5 SEM-EDX spectrum of MMT (a, b) and QCD-MMT (c, d) sediments

et al. 2016). Moreover, a probable mechanism was sites than −OH. The ion exchange could take place
proposed to be ion exchange between chromium through the following reactions (Wang et al.
and surface −OH groups or hydroxide groups of 2016):
Al polycations proceeded in the interlayer spaces.
≡S−OH− þ Hþ →≡S−OHþ
At lower pH, −OH groups were protonized to 2

−OH2+, which facilitates the ligand exchange since


–OH2+ is easier to displace from metal binding
≡S−OH− þ HCrO−4 →≡S−HCrO−4 þ OH−
Table 3 Composition of MMT and QCD-MMT determined by
SEM-EDX
≡S−OHþ − −
2 þ HCrO4 →≡S−HCrO4 þ H2 O
Element MMT QCD-MMT

Wt.% At.% Wt.% At.%


For methyl orange, the adsorption capacities
CK 02.69 04.73 10.27 17.23 changed from 12.43 to 23.72 mg/g at different pH
OK 39.16 51.67 36.19 45.58 values. With pH value of 3 ~ 11, the adsorption
NaK 02.70 02.48 01.66 01.45 capacities were around 22 mg/g without significant
MgK 02.91 02.53 02.39 01.98 changes. The hydroxyl of QCD-MMT might form
AlK 10.98 08.59 09.26 06.92 hydrogen bonds with N atoms of methyl orange
SiK 35.72 26.85 30.48 21.86
skeleton. Meanwhile, hydroxyl proton can produce
ClK 00.09 00.05 03.14 01.78
electrostatic adsorption with SO3− contained in MO
molecules (Yan et al. 2016). Moreover, the inclusion
278 Page 10 of 17 Water Air Soil Pollut (2017) 228:278

Fig. 6 TGA curves of MMT and


QCD-MMT

between methyl orange and β-CD unit of QCD- combination of host-guest inclusion, electrostatic
MMT also benefit adsorption. When pH was of 13, repulsion, and hydrogen bond interaction (Li
the hydrogen bonds might not be easy to form be- et al. 2011; Shen et al. 2015). When pH was
tween β-CD unit and methyl orange, since the OH− low, both β-CD unit and p-nitrophenol become
in the solution would compete with methyl orange protonated, so p-nitrophenol cannot be adsorbed
and form hydrogen bonds with β-CD unit. This on β-CD efficiently for the electrostatic repulsion
would further decrease electrostatic adsorption and between them. This results in lower adsorption
then inclusion, resulting in lower adsorption (Yan capacity at pH of 1 and 3. When pH value in-
et al. 2016). creased, protonation of β-CD and p-nitrophenol
For p-nitrophenol, the maximum adsorption was became weak. This might benefit the host-guest
obtained at pH of 5, which was 21.45 mg/g. Also, inclusion between them, and reach the maximum
at pH of 11, the adsorption was high (19.36 mg/g) adsorption at pH of 5. As the pH became higher
as well. The fluctuation could be explained by the (pHs of 7 and 9), inclining adsorption was

Fig. 7 DTA curves of MMT and


QCD-MMT
Water Air Soil Pollut (2017) 228:278 Page 11 of 17 278

Table 4 TGA-DTG results of MMT and QCD-MMT

Sample Temperature of Residue at 800 °C Organic content CD content QCD content Total EPTAC content
maximum (wt.%) (wt.%) (wt.%) (wt.%) (wt.%)
decomposition
rate (°C)

T1 T2 T3

MMT 64 – 661 85.08 / / / /


QCD- 60 304 635 75.69 16.25a 8.03b 9.10b 8.22b
MMT 92
a
Organic content calculated from Eq. (1)
b
CD content calculated from following equations:
MQCD * nQCD + MEPTAC * nEPTAC = organic content
nQCD + nEPTAC = mN / MN
nQCD = nCD
where MQCD stands for molar mass of QCD, MEPTAC stands for molar mass of EPTAC, nQCD stands for mole number of QCD, nEPTAC stands
for mole number of EPTAC, mN stands for mass of N element, and MN stands for molar mass of N element

observe. When the pH reached pH of 11, the adsorption capacity could be attributed to host-
alkaline environment greatly weakened the proton- guest inclusion and hydrogen bond interaction.
ation of β-CD and p-nitrophenol, so the increased However, when the pH was of 13, certain amount

Fig. 8 Influence of usage on QCD-MMT adsorption capacity and


adsorbates’ equilibrium concentration (contact time 24 h; initial Fig. 9 Influence of adsorbates’ initial concentrations on adsorp-
concentration Cr(VI) 120 mg/L, methyl orange 50 mg/L, p- tion capacity on different temperatures (contact time 24 h; QCD-
nitrophenol 50 mg/L; temperature 30 °C; neutral) MMT usage 0.05 g; neutral)
278 Page 12 of 17 Water Air Soil Pollut (2017) 228:278

Fig. 10 1 QCD-MMT’s adsorption capacity was plotted as a b Methyl orange, QCD-MMT 0.05 g, 30 °C, neutral, initial con-
function of time. Adsorption kinetics was investigated using 2 centration 250 mg/L. c p-Nitrophenol, QCD-MMT 0.05 g, 30 °C,
pseudo-first-order and 3 pseudo-second-order models. a Cr(VI), neutral, initial concentration 25 mg/L
QCD-MMT 0.05 g, 30 °C, neutral, initial concentration 120 mg/L.

of p-nitrophenol might form p-nitrophenol sodium for the data, considering higher R2 and higher agree-
salt with sodium hydroxide. The hydrophobic in- ment between the calculated and experimental qe
clusion could be impacted, causing decreasing values. This explained chemisorption controls the
adsorption. rate of adsorption. Table 5 and Fig. 12 showed the
Overall, the adsorption was pH dependent. More detailed values. In chemisorption, the adsorbate ad-
investigation would be conducted in the future to ex- hered on QCD-MMT by formation of a chemical
plore more details about the mechanism. bond, and also searched for adsorption sites. This
would maximize their coordination number with the
3.3 Adsorption Kinetics surface (Kumar and Kirthika, 2009; Mishra et al.
2017). Furthermore, the kinetics of p-nitrophenol
Pseudo-first-order and pseudo-second-order models and methyl orange was faster than that of Cr(VI),
were used to evaluate the sorption kinetic. The ki- suggesting that the β-CD cavities inclusion process
netic parameters for adsorption of three different was faster than MMT complexing, comparing
samples were given in Table 5. For all three sam- pseudo-second-order rate constant k2 values (Zhao,
ples, the pseudo-second-order equation fitted better Repo, Yin, et al., 2015a).
Water Air Soil Pollut (2017) 228:278 Page 13 of 17 278

monolayer coverage of methyl orange or p-


nitrophenol molecule was formed at the outer sur-
face of QCD-MMT (Malik 2004). Since β-CD was
the effective component for organic methyl orange
and p-nitrophenol, it would be inferred that the β-
CD distributed homogenously at the final QCD-
MMT composite. Debnath et al. reported the sim-
ilar observation (Debnath et al. 2017). They used
β-CD modified pinecone to remove safranine O,
brilliant green, and methylene blue; the adsorption
also followed the Langmuir isotherm.
The introduction of β-CD to the QCD-MMT com-
posite enabled the cavity structure to capture organic
molecules. It may be inferred that the interactions be-
tween the organic molecules used in this article and the
β-CD were primarily due to host-guest inclusion. Also,
methyl orange and p-nitrophenol molecules might have
possible hydrogen bonding with the hydroxyl groups on
the surface of the β-CD.
Cr(VI) adsorption fits Freundlich isotherm mod-
Fig. 11 Influence of pH on QCD-MMT adsorption capacity and
el better, suggesting a multilayer adsorption for
adsorbates’ equilibrium concentration (contact time 24 h; initial
concentration Cr(VI) 120 mg/L, methyl orange 50 mg/L, p- Cr(VI). This result could be attributable to the
nitrophenol 50 mg/L; temperature 30 °C; neutral) large montmorillonite spacing and the cationic ad-
sorption sites for Cr(VI) due to surfactant loadings.
3.4 Adsorption Isotherms These cationic adsorption sites might be favorable
to the Cr(VI) adsorption through anion exchange
As shown in Fig. 12 and Table 6, the Langmuir between the counter anion of surfactant and
isotherm model provides the better fit for the HCrO 4 − or Cr 2 O 7 2− (Wang et al. 2016). The
methyl orange and p-nitrophenol equilibrium data. Cr(VI) adsorption on QCD-MMT might also
This suggested that the adsorption for methyl or- depended on electrostatic attraction and the ion
ange and p-nitrophenol was monolayer, or homo- exchange between chromium and surface −OH
geneous adsorption (Debnath et al. 2017). Each groups (Li et al. 2014). Also, the value of n from
methyl orange or p-nitrophenol molecule might Freundlich isotherm model was found to be slight-
have equal adsorption affinity towards QCD- ly higher than 1, which means that the adsorption
MMT (Shen et al. 2015). It was possible that the for Cr(VI) was a linear sorption.

Table 5 Adsorption kinetics of methyl orange on QCD-MMT

Kinetic models Equation Adsorbates Parameters

k Experimental Calculated R2
qe (mg/g) qe (mg/g)

Pseudo-first order qt ¼ qe1 −qe1 e−k 1 t Cr(VI) 0.068 43.000 41.943 0.8819
Methyl orange 0.072 50.000 48.227 0.8675
p-Nitrophenol 0.045 6.500 6.091 0.8318
Pseudo-second order t
qt ¼ qt þ k 21q2 Cr(VI) 0.002 43.000 45.558 0.9992
e2 e2
Methyl orange 0.004 50.000 49.383 0.9884
p-Nitrophenol 0.007 6.500 7.136 0.9931
278 Page 14 of 17 Water Air Soil Pollut (2017) 228:278

Fig. 12 Adsorption isotherm on QCD-MMT using L Langmuir and F Freundlich isotherm model. (The solid line represented the fit curves.
Neutral solutions; QCD-MMT usage of 0.05 g; contact time of 24 h a Cr(VI); b methyl orange; c p-nitrophenol)

3.5 Adsorption Thermodynamics 4 Conclusions

As shown in Table 7, for all three adsorbates, the posi- In this paper, the adsorption capacities of QCD-MMT
tive standard enthalpy change ΔH0 values indicated that composite towards Cr(VI), methyl orange, and p-
the adsorption was endothermic process. The negative nitrophenol were studied. The adsorption capacities
standard free energy change ΔG0 values and positive might be attributed both to the hydrophobic cavity of
standard entropy change ΔS0 values indicated that the β-CD and the special metal-removing capacities of
adsorption process was spontaneous with high affinity. MMT.
ΔG0 values decreased with the increasing temperature, β-CD was cyclic oligosaccharides including glucose
implying more efficient adsorption at higher tempera- units linked by α-1,4-glucosidic bonds (Zhang et al.
tures. For Cr(VI) adsorption, the ΔG0 values became 2013), which has a hydrophobic inner cavity and a
negative when temperature was higher, around 50 °C. hydrophilic exterior (Qiao et al. 2010). This special
Water Air Soil Pollut (2017) 228:278 Page 15 of 17 278

Table 6 Isotherm parameter for adsorption of Cr(VI), methyl orange, and p-nitrophenol on QCD-MMT

Temperature Langmuir Freundlich

R2 qmax KL R2 KF n
K mg/g L/mg (mg/g)(mg/L)n

Cr(VI) 303.15 0.096 980.39 0.001 0.991 0.98 1.038


313.15 0.426 423.73 0.002 0.976 1.19 1.079
323.15 0.963 159.24 0.007 0.981 1.93 1.319
Methyl orange 303.15 0.994 75.47 0.016 0.963 2.59 1.597
313.15 0.995 87.11 0.866 0.978 2.44 1.526
323.15 0.988 102.77 0.011 0.99 2.39 1.468
p-Nitrophenol 303.15 0.988 44.05 0.126 0.73 8.74 2.927
313.15 0.986 47.12 0.109 0.814 8.84 2.849
323.15 0.979 50.15 0.105 0.811 9.00 2.773

cavity structure of β-CD constituted a kind of non-polar The composite also utilized MMT’s significant
micro-environment, allowing a number of small object adsorption capacities towards metal ions. This study
organic molecules and inorganic ions to enter. It could used Cr(VI) as a representative. The adsorption of
package multiple object molecules to form supramolec- Cr(VI) was highly dependent on adsorbent dose,
ular complexes. The adsorption of methyl orange and p- initial concentration, temperature, and contact time.
nitrophenol in this study could be ascribed to the inclu- The adsorption kinetics of QCD-MMT followed the
sion in the hydrophobic cavity of β-CD and hydrogen pseudo-second-order model. Adsorption of Cr(VI)
bond interaction (Badruddoza et al. 2013; Shen et al. also fits better in the Freundlich model, inferring a
2015). The kinetics study also showed that p- multilayer adsorption.
nitrophenol and methyl orange adsorption followed Thermodynamic analysis showed that the adsorp-
pseudo-second-order model and was faster than that tion was all endothermic process and could be spon-
of Cr(VI), suggesting that the β-CD cavities inclu- taneous at given temperature range, except for
sion process was faster than MMT complexing Cr(VI), of which adsorption should be at much
(Zhao, Repo, Yin, et al., 2015a). Moreover, the higher temperature.
isotherms fitting results of methyl orange and p- Further studies would be focused more on the
nitrophenol adsorption indicated that the absorption adsorption towards mixed wastewater containing
of these two molecules were better described by different types of pollutants. More detailed inves-
Langmuir model, which was monolayer, or tigation into adsorption mechanism would be con-
homogeneous. ducted as well.

Table 7 Thermodynamic parameters of Cr(VI), methyl orange, and p-nitrophenol adsorption

Adsorbates ΔH0, kJ/mol ΔS0, J/(mol k) ΔG0, kJ/mol ΔG0, kJ/mol ΔG0, kJ/mol
303.16 K 313.16 K 323.16 K

Cr(VI) 5.38 17.03 0.22 0.05 −0.12


Methyl orange 2.48 8.43 −0.08 −0.16 −0.24
p-Nitrophenol 4.71 21.83 −1.90 −2.12 −2.34
278 Page 16 of 17 Water Air Soil Pollut (2017) 228:278

Acknowledgements The Science Foundations from the Depart- Shen, H. M., Zhu, G. Y., Yu, W. B., Wu, H. K., Ji, H. B., Shi, H. X.,
ment of Education, Fujian Province (JA14428) and from Liming She, Y. B., & Zheng, Y. F. (2015). Fast adsorption of p-
Vocation University (LZ2016109) were acknowledged. nitrophenol from aqueous solution using β-cyclodextrin
grafted silica gel. Applied Surface Science, 356, 1155–1167.
Sowmya, A., & Meenakshi, S. (2013). An efficient and regenera-
ble quaternary amine modified chitosan beads for the remov-
al of nitrate and phosphate anions. Journal of Environmental
References Chemical Engineering, 1(4), 906–915.
Vašák, M., & Meloni, G. (2011). Chemistry and biology of mam-
malian metallothioneins. JBIC Journal of Biological
Badruddoza, A. Z. M., Shawon, Z. B. Z., Wei, J. D. T., Inorganic Chemistry, 16(7), 1067–1078.
Hidajat, K., & Uddin, M. S. (2013). Fe3O4/cyclodex- Wang, G., Hua, Y., Su, X., Komarneni, S., Ma, S., & Wang, Y.
trin polymer nanocomposites for selective heavy metals (2016). Cr(VI) adsorption by montmorillonite nanocompos-
removal from industrial wastewater. Carbohydrate ites. Applied Clay Science, 124–125, 111–118.
Polymers, 91(1), 322–332.
Xie, W., Xie, R., Pan, W., Hunter, D., Koene, B., Tan, L., & Vaia,
Chen, J., Qiu, F., Xu, W., Cao, S., & Zhu, H. (2015). Recent R. (2014). Thermal stability of quaternary phosphonium
progress in enhancing photocatalytic efficiency of TiO2- modified montmorillonites. Chemistry of Materials, 14(11),
based materials. Applied Catalysis A General, 495, 131–140. 4837–4845.
Debnath, S., Ballav, N., Maity, A., & Pillay, K. (2017). Xing, M., & Wang, J. (2016). Nanoscaled zero valent iron/
Competitive adsorption of ternary dye mixture using pine graphene composite as an efficient adsorbent for Co(II) re-
cone powder modified with β-cyclodextrin. Journal of moval from aqueous solution. Journal of Colloid & Interface
Molecular Liquids, 225, 679–688. Science, 474, 119–128.
Ezquerro, C. S., Ric, G. I., Miñana, C. C., & Bermejo, J. S. (2015). Yan, J., Zhu, Y., Qiu, F., Zhao, H., Yang, D., Wang, J., & Wen, W.
Characterization of montmorillonites modified with organic (2016). Kinetic, isotherm and thermodynamic studies for
divalent phosphonium cations. Applied Clay Science, 111, 1– removal of methyl orange using a novel β-cyclodextrin func-
9. tionalized graphene oxide-isophorone diisocyanate compos-
Kumar, P. S., & Kirthika, K. (2009). Equilibrium and kinetic study ites. Chemical Engineering Research & Design, 106, 168–
of adsorption of nickel from aqueous solution onto bael tree 177.
leaf powder. Journal of Engineering Science & Technology, Yao, W., Rao, P., Lo, I. M., Zhang, W., & Zheng, W. (2014).
4(4), 351–363. Preparation of cross-linked magnetic chitosan with quaterna-
Li, J., Chen, C., Zhao, Y., Hu, J., Shao, D., & Wang, X. (2013). ry ammonium and its application for Cr(VI) and P(V) remov-
Synthesis of water-dispersible Fe3O4 @β-cyclodextrin by al. Journal of Environmental Sciences, 26(12), 2379–2386.
plasma-induced grafting technique for pollutant treatment. Yuan, C., Liu, B., & Liu, H. (2015). Characterization of
Chemical Engineering Journal, 229, 296–303. hydroxypropyl-β-cyclodextrins with different substitution
Li, T., Shen, J., Huang, S., Li, N., & Ye, M. (2014). Hydrothermal patterns via FTIR, GC-MS, and TG-DTA. Carbohydrate
carbonization synthesis of a novel montmorillonite supported Polymers, 118, 36–40.
carbon nanosphere adsorbent for removal of Cr (VI) from Zhang, W., Chen, M., Gong, X., & Diao, G. (2013). Universal
waste water. Applied Clay Science, 93–94(3), 48–55. water-soluble cyclodextrin polymer–carbon nanomaterials
Li, X., Zhao, B., Zhu, K., & Hao, X. (2011). Removal of with supramolecular recognition. Carbon, 61(5), 154–163.
nitrophenols by adsorption using β-cyclodextrin modified Zhao, D., Zhao, L., Zhu, C., Tian, Z., & Shen, X. (2009). Synthesis
zeolites. Chinese Journal of Chemical Engineering, 19(6), and properties of water-insoluble β-cyclodextrin polymer
938–943. crosslinked by citric acid with PEG-400 as modifier.
Liu, Y., Li, M., & He, C. (2016). Removal of Cr(VI) and Hg(II) Carbohydrate Polymers, 78(1), 125–130.
ions from wastewater by novel β-CD/MGO-SO3H compos- Zhao, F., Repo, E., Meng, Y., Wang, X., Yin, D., & Sillanpää, M.
ite. Colloids & Surfaces A Physicochemical & Engineering (2015b). An EDTA-β-cyclodextrin material for the adsorp-
Aspects, 512, 129–136. tion of rare earth elements and its application in
Malik, P. K. (2004). Dye removal from wastewater using activated preconcentration of rare earth elements in seawater. Journal
carbon developed from sawdust: Adsorption equilibrium and of Colloid & Interface Science, 465, 215–224.
kinetics. Journal of Hazardous Materials, 113(1–3), 81–88. Zhao, F., Repo, E., Sillanpää, M., Meng, Y., Yin, D., & Tang, W. Z.
Mishra, S. R., Chandra, R., Jipsi, K. A., & Savariya, D. B. (2017). (2015c). Green synthesis of magnetic EDTA-and/or DTPA-
Kinetics and isotherm studies for the adsorption of metal ions cross-linked chitosan adsorbents for highly efficient removal
onto two soil types. Environmental Technology & of metals. Industrial & Engineering Chemistry Research,
Innovation, 7, 87–101. 54(4), 1271–1281.
Qiao, J., Zhang, H. Y., & Yu, L. (2010). Synthesis of β- Zhao, F., Repo, E., Yin, D., Meng, Y., Jafari, S., & Sillanpää, M.
cyclodextrin-modified carbon nanocrystals and their fluores- (2015a). EDTA-cross-linked β-cyclodextrin: An environ-
cent behavior. Science Bulletin, 55(25), 2835–2839. mentally friendly bifunctional adsorbent for simultaneous
Rosa, S., Laranjeira, M. C., Riela, H. G., & Fávere, V. T. (2008). adsorption of metals and cationic dyes. Environmental
Cross-linked quaternary chitosan as an adsorbent for the Science & Technology, 49(17), 10570–10580.
removal of the reactive dye from aqueous solutions. Zhao, F., Repo, E., Yin, D., & Sillanpää, M. E. (2013). Adsorption
Journal of Hazardous Materials, 155(1–2), 253–260. of Cd(II) and Pb(II) by a novel EGTA-modified chitosan
Water Air Soil Pollut (2017) 228:278 Page 17 of 17 278

material: Kinetics and isotherms. Journal of Colloid & polyacrylamide microcomposite adsorbent. Journal of Water
Interface Science, 409(11), 174–182. Process Engineering, 4(4), 47–57.
Zhao, F., Tang, W. Z., Zhao, D., Meng, Y., Yin, D., & Sillanpää, Zheng, W. (2001). Toxicology of choroid plexus: Special refer-
M. (2014). Adsorption kinetics, isotherms and mechanisms ence to metal-induced neurotoxicities. Microscopy Research
of Cd(II), Pb(II), Co(II) and Ni(II) by a modified magnetic & Technique, 52(1), 89–103.
Water, Air & Soil Pollution is a copyright of Springer, 2017. All Rights Reserved.

You might also like